Download 1 Protonolysis of Fe-C bonds of a Diiminopyridineiron(II) Dialkyl

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Nucleic acid analogue wikipedia , lookup

Metabolism wikipedia , lookup

Isotopic labeling wikipedia , lookup

NADH:ubiquinone oxidoreductase (H+-translocating) wikipedia , lookup

Siderophore wikipedia , lookup

Multi-state modeling of biomolecules wikipedia , lookup

Biochemistry wikipedia , lookup

Hepoxilin wikipedia , lookup

Ligand binding assay wikipedia , lookup

Nuclear magnetic resonance spectroscopy of proteins wikipedia , lookup

Evolution of metal ions in biological systems wikipedia , lookup

Metalloprotein wikipedia , lookup

Transcript
Protonolysis of Fe-C bonds of a Diiminopyridineiron(II) Dialkyl Complex by Acids of Different Strengths:
Influence of Monoanionic Ligands on the Spectroscopic Properties of Diiminopyiridine-FeY2 Complexes.
*
M. Ángeles Cartes, Pilar Palma, John J. Sandoval, Juan Cámpora and Eleuterio Álvarez.
Instituto de Investigaciones Químicas, CSIC – Universidad de Sevilla. c/ Américo Vespucio, 49, 41092, Sevilla,
Spain.
Summary
The
reaction
of
the
dialkyliron
complex
[Fe(CH2SiMe3)2(
Mes
BIP)]
(
Mes
BIP
=
2,6-bis((N-
mesityl)acetimidoyl)pyridine) with protic acids (HY) of different strengths (Y = C6F5O, CF3CO2, Cl, CF3SO3)
invariably leads to the cleavage of both Fe-C bonds, independently of the Fe/HY ratio used (either 1:2 or 1:1),
affording the corresponding complexes [FeY2(
Mes
BIP)]. Relevant spectroscopic features of these compounds, such
1
as paramagnetic H NMR shifts and UV-VIS absorption bands, exhibit a marked dependence on the nature of Y.
Keywords: 2,6-bisiminopyridine ligands, iron, alkyl complexes, olefin polymerization.
1. Introduction
Olefin polymerization or oligomerization catalysts based on iron complexes of 2,6-bisiminopyridine (BIP)
ligands have attracted much interest due to their high activity and the abundance and low toxicity and of iron.
[1]
In
addition, the modular design of BIP ligands facilitates the variation of the stereoelectronic environment of the active
centre, enabling a precise control of the molecular weight of the polymers. It is usually assumed that, similarly to
other polymerization systems, activation of [FeX2(BIP)] complexes with alumoxanes or other organoaluminumbased co-catalysts give rise to catalytically active alkyliron species. Such a classic Ziegler-Natta mechanism gained
iPr
+
iPr
strong support in 2005, when Chirik prepared cationic complexes of the type [Fe(R)(S)( BIP)] ( BIP= BIP ligand
with 2,6-diisopropylphenyl as aryl substituent in the imine; S = OEt2, THF, or none) by protonation of dialkyl
+
- [2]
precursors with [HNPhMe2 ] [BPh4] ,
and demonstrated that such cationic iron alkyls behave as highly active
single-component catalysts for ethylene polymerization. However, the precise nature of the active species on the
real catalysts generated with the aid of alkylaluminum co-catalysts is still the subject of some controversy. The
latter are known to play a very important role in the catalytic process, influencing both the activity and the molecular
*
E-mail address: [email protected] (J. Cámpora).
1
weight distribution of the polyolefinic products.
[3]
Indeed, spectroscopic investigations of the aluminium-activated
iron catalysts by Bryliakov and Talsi have revealed that the interaction of [FeX2(BIP)] complexes with
organoaluminum reagents gives rise to both neutral and cationic bimetallic Fe/Al species that very likely have an
active participation in the polymerization process.
[4]
In addition, it has been recognized that the counteranion that
balances the electric charge of active cationic species plays a role of crucial importance in the performance of most
homogeneous Ziegler-Natta polymerization catalysts.
[5]
Thus, it seems very likely that iron complexes of the type
[Fe(R)(Y)(BIP)] (where Y symbolizes an anionic ligand) should exhibit significant differences in their ability to act as
polymerization catalysts, as the coordinating strength of Y or its ability to interact with the co-catalysts can be
varied widely. Several years ago, we reported a general methodology that provides access to iron dialkyl
complexes of the type [Fe(CH2SiMe3)2(BIP)],
[6]
and we wondered whether these complexes could react selectively
with protic acids HY of different strengths to afford the desired [Fe(CH2SiMe3)(Y)(BIP)] complexes. As we show in
this contribution, it turned out that such mixed ligand compounds are not stable or cannot be produced through this
route. Instead, the protonation reaction affords symmetrical [FeY2(BIP)] derivatives. This allowed us to compare
some of the key spectroscopic features of these compounds and analyse how these properties are influenced by
the nature of the anionic Y ligand.
2. Results and Discussion.
We investigated the reactions of the readily available dialkyl complex [Fe(CH2SiMe3)2(
Mes
BIP)] (1) with four
protic acids of different strengths: Pentafluorophenol (pKa = 5.4), trifluoroacetic acid (pKa = 0.2), hydrogen chloride
(pKa = -7) and triflic acid (pKa = -12).
[7,8]
Stoichiometric (1:1) amounts of the acids diluted in THF were slowly added
to the solutions of the iron dialkyl in the same solvent at -80 ºC, and then allowed to slowly warm to room
temperature. In spite of the care taken to control the reaction conditions, these reactions invariably led to products
2 – 5 resulting from the cleavage of both Fe-C bonds of 1 (Scheme 1). The products were precipitated by addition
of hexane, leaving purple mother liquors containing unreacted 1. As expected, higher yields of all four products
were obtained when the acids and the iron alkyl were reacted in 2:1 ratio. These results suggest that the nonsymmetrical alkyl complexes [Fe(CH2SiMe3)(Y)(
Mes
BIP)] are unstable and rapidly disproportionate in solution
affording a mixture of the corresponding symmetrical complexes. This conclusion is also supported by Chirik’s
observation that the reaction of the mixed complex [Fe(CH2SiMe3)(Cl)(Py)2] with
iPr
iPr
BIP does not afford the intended
(chloro)trimethylsilylmethyl derivative [Fe(CH2SiMe3)2(Cl)( BIP)], but a mixture of the dialkyl and dichloro
complexes.
[9]
2
N
N Fe
N
Y
X
SiMe 3
N
N
Fe
Me 3Si
HY
1:1
N
SiMe 3
1
N
or
N Fe
N
N Y
N
N Fe
H 2O
Y
Y
Y
2, Y = C6F 5O
3, Y = CF 3CO2
4, Y = Cl
5, Y = CF 3SO 3
Scheme 1
Apart from the well-known chloro derivative 4,
[3]
none of the rest of the products, 2. 3 and 5, has been described
previously. They are all paramagnetic with µeff = 5.0 – 5.6 µB at room temperature, consistent with a high-spin
configuration with four unpaired electrons. Crystals of 2 and 5 suitable for X-ray analysis were obtained by
recrystallization. Figures 1 and 2 show ORTEP views of these two complexes, and Table 1 collects selected bond
lengths and angles. Remarkably, very few iron bisiminopyridine complexes containing alkoxo or aryloxo ligands
have been reported before,
[10]
and to the best of our knowledge, no dialkoxo or diaryloxo derivatives have been
reported before. Similarly to the analogous halide complexes [FeX2(
Mes
BIP)] (X = Cl, Br), the iron centre of the
aryloxide 2 exhibits a distorted trigonal bipyramidal geometry, with the imine nitrogen atoms occupying the axial
positions. The τ parameter,
[11]
that describes the distortion degree between perfect bipyramidal trigonal (τ =1) and
square pyramidal (τ =0) geometries, takes the value 0.84 for this compound. A crystallographically imposed mirror
plane bisects the molecule through the iron and the three nitrogen atoms and relates the pentafluroroaryloxide
moieties. The latter are oriented in such a way that one of the ortho fluorine substituents (F5) approaches to the
iron atom. This conformation could be favoured by an attractive electrostatic interaction, but the Fe-F5 distance
(3.0149(12) Å) is too long to mean any significant chemical bonding. As it is usually found in this type of
compounds, the Fe-N bonds involving the imino groups are somewhat different, decreasing the overall molecular
symmetry from C2v (if both Fe-imine bonds were identical) to Cs. The structure of complex 5 contains, in addition to
the triflate ligands, a molecule of water attached to the iron centre. The presence of an aqua ligand is also revealed
by a strong absorption at 3345 cm
-1
in the IR spectrum of this compound. Very likely, this water comes from
adventitious traces of moisture in the solvents during the synthesis or the recrystallization of this complex, and its
presence reveals the strong Lewis acidity of the corresponding bis-triflate precursor. Britovsek has reported related
3
Figure 2. ORTEP view of the X ray structure of
compound 5.
Figure 1. ORTEP view of the X ray structure of
compound 2
Table 1: Selected bond distances (Å) and bond angles (º) for compounds 2 and 5.
Distances (Å)
2
5
Angles (deg)
2
5
Fe-N1
2.0943(17)
2.089(2)
O1-Fe-O1’/O4
97.06(7)
84.32(7)
Fe-N2
2.2923(17)
2.236(2)
N1-Fe-N3
73.97(6)
74.50(8)
Fe-N3
2.2368(17)
2.210(2)
N2-Fe-N3
147.58(6)
147.71(8)
Fe-O1
1.9840(11)
2.1084(18)
O1-Fe-N1
131.36(3)
62.26(7)
Fe-O4
--
2.2277(19)
O4-Fe-N1
--
93.82(7)
Fe-O7
--
2.1288(18)
O7-Fe-N1
--
110.88(8)
O7···(H)···O3
--
2.804(3)
O1-Fe-N2
103.14(5)
118.47(7)
O1-Fe-N3
98.14(5)
93.82(7)
Fe(II) and Mn(II) bis-triflato complexes with BIP ligands. Interestingly, the Mn(II) derivative containing the
Me2
BIP
ligand
(N-aryl
[Mn(OTf)2(OH2)(
Me2
iPr
BIP)].
[Fe(OTf)(SPh)( BIP)].
groups
[12]
[13]
=
2,6-dimethylphenyl)
was
also
isolated
as
the
monohydrate
Goldberg has recently reported a mixed iron(II) thiolate-triflate complex.
The geometry of the iron centre in 5 is approximately octahedral, with the aqua
ligand and one of the triflate ligands occupying “axial” positions, i. e., perpendicular to the main coordination
plane defined by the three N atoms of the BIP ligand and the second triflate group. This configuration
contrasts with that preferred by the analogous manganese complex [Mn(OTf)2(OH2)(
Me2
BIP)], which shows
both triflate groups in the axis and the aqua ligand sharing the equatorial plane with the BIP ligand.
[9]
A
hydrogen bond links the aqua ligand of 5 to the cis triflate group. The distance between the oxygen atoms
4
involved in this interaction, O3 and O7, (2.804(3) Å) suggests that the H bridge is relatively weak, as it is
longer than those observed in typical O···H···O hydrogen bonds, for instance those formed in water or in
carboxylic acids (2.6 – 2.7 Å).
[14]
The Fe-O bonds (either those involving the triflate or aqua ligands) are 2.1 –
2.2 Å long, significantly longer than the Fe-O bonds in the aryloxide 2 (1.9840(11) Å). This difference is due
in part to the higher coordination number in 5, but they also reflect the weaker nature of the Fe-O bonds
involving the aqua and triflate anions, the latter predominantly ionic in character. Although we were unable to
grow X-ray quality crystals of carboxylate 3, this compound probably has a pentacoordinated structure with
terminally bound carboxylate ligands, as observed for related carboxylate compounds containing terpy
related tridentate N,N,N ligands.
[16]
[15]
or
Terminal coordination of the trifluoroacetate ligands is supported by the
large separation between the IR absorptions for the νs and νas modes of the carboxylate group, observed at
-1
-1 [17]
1697 and 1260 cm (Δν = 437 cm ).
The colours of complexes 1 – 5 are strikingly varied. While the dialkyl 1 is purple, perfluorophenolate
2 is green, the trifluoroacetate and triflate derivatives 3 and 5 are burgundy-red and magenta, respectively,
and the chloro complex 4 is dark blue. The origin of these colours is a broad absorption band in the visible
spectrum, whose position varies depending on the anionic ligands coordinated to the iron centre (Figure 3).
The absorption maximum (λmax) appears at 550 nm for 1 and shifts to longer wavelengths for the products
arising from the protonation with acid, in the order 1 < 2 (580 nm) < 3 (625 nm) < 4 (680 nm). This band has
also been observed in the visible spectra of other iron-BIP complexes and assigned to a charge transfer
transition involving the BIP ligand.
[18]
The observed trend indicates that the energy gap between the orbitals
responsible for the transition decreases as the acid HY from which the complex arises becomes stronger, or
-
what is the same, as the anion Y becomes more electronegative. This points to a ligand to metal (LMCT)
rather than a metal to ligand charge transfer band (MLCT), since electron withdrawing from the metal is
expected to lower the energy of metal-centred orbitals, while leaving those of the BIP ligand relatively
unaffected. Triflate complex 5 is the exception to the mentioned trend, as its visible absorption occurs at 550
nm, when it would be expected to be the lowest in energy for the series. This apparent exception is readily
explained by the presence of the additional aqua ligand, which contributes to compensate the electronic
deficiency caused by the strongly electron-withdrawing triflate ligands on the metal centre.
5
Figure 3. UV-VIS spectra of complexes 1 – 5.
1
High spin Fe(II) complexes with BIP ligands give rise to useful NMR spectra. Typically, the H signals
have linewidths between a few and several hundred hertz and can be readily observed, except those
belonging to groups directly attached to the metal centre. These signals can be assigned on the basis of
their intensity, width and their characteristically large chemical shifts. Recognizing trends in the spectra of
series of related complexes is highly helpful in the task of assigning resonances, and constitutes a useful tool
1
for the identification of new complexes. The H NMR spectra of complexes 2 – 5 provide some of such useful
trends (see Figure 4). Since none of the ligands Y contain H atoms, these spectra show only signals
corresponding to the
Mes
BIP ligand. They all exhibit the same number of signals, corresponding to apparent
C2v molecular symmetry. In the case of the triflate complex, the simplicity of the spectrum is not consistent
with the lower symmetry observed in the solid state, indicating that in solution the aquo land triflate ligands
exchange their relative positions, probably via a mechanism involving OTf dissociation. This is supported by
the
19
F spectrum, which shows a single resonance at δ -14.6 ppm in dichloromethane, which shifts to -71.0
ppm in acetonitrile, close to the position expected for an uncoordinated triflate anion.
[12]
One of the most
evident features of these spectra is the strong sensitivity of the chemical shift of some of the
Mes
BIP signals
to the nature of ligands Y. The signal H4 is particularly noteworthy, as it shifts to higher field in the order Y =
Cl (complex 4, 34.8 ppm) ≈ C6F5O (2, 35.5 ppm) > triflate (5, 22.7 ppm) > trifluoroacetate (3, -11.2 ppm).
The signal for the α -methyl of the imino group shows the same tendency, but with the opposite sign: 2 (-22.6
ppm) ≈ 4 (-22.2) < 5 (-3.6 ppm) < 3 (+4.1 ppm). The inverse correlation between the chemical shifts of the
H4 and α-Me signals holds for the dialkyl 1, for which they take extreme values, +284.0 and -184.3,
respectively. The analogous signals in other Fe(II) dialkyl complexes containing BIP ligands exhibit similar
extreme paramagnetic shifts.
[6,9,10]
The observed H4/α-Me sequences bear no simple relationship with the
6
donor capacity of Y, since paramagnetic chemical shifts depend on the spin density on the observed nucleus
rather than on partial electric charge effects. However, the much larger paramagnetic shifts of the position of
H4 and α-Me signals in 1 enables us to discriminate between simple coordination complexes containing mild
σ-donor Y groups from organometallic species.
1
Figure 4. H NMR spectra of complexes 2- 5. Residual peaks of deuterated solvent (C6D6 for 2 and CD2Cl2
for 3, 4 and 5) are marked with asterisk (*).
3. Conclusions
Our attempts to selectively cleave one Fe-C bonds of the dialkyl 1 with a variety of protic acids HY
of different strengths have been unsuccessful, as these reactions invariably lead to [FeY2(
Mes
BIP)]
complexes arising from the cleavage of both Fe-C bonds. We evaluated the influence of ligands Y on the
spectroscopic properties of this series of complexes. The charge transfer absorption band responsible for the
colour of these complexes shifts to longer wavelengths as the electron density on the Fe atom, controlled by
1
the Y ligands, decreases. Although the chemical shift of some H NMR signals of the
Mes
BIP ligand (in
particular, those due to pyridine H4 and the imine α-Me group) are very sensitive to the nature of Y, no clearcut relationship is observed. These exhibit much larger paramagnetic shifts in the organometallic parent
compound 1 than in the coordination complexes 2 – 5.
7
4. Experimental Section
4.1. General considerations. All manipulations were carried out under inert atmosphere by using
conventional Schlenk techniques or a nitrogen-filled glovebox. Solvents were rigorously dried and degassed
prior to use. Microanalysis were performed by the Analytical Service of the Instituto de Investigaciones
Químicas. IR spectra were recorded with Bruker Vector 22 or Tensor 27 spectrometers, and UV-VIS spectra
with a Perkin-Elmer Lambda 12 spectrophotometer, using special cuvettes provided with a gastight Young
®
Teflon valve. NMR spectra were recorded with Bruker 300 and 400 MHz spectrometers. Chemical shifts are
1
expressed relative to TMS. The H NMR residual resonance of the solvent was used as internal standard
and an external reference of CF3CO2H was used for the
19
F spectra of 5. Magnetic susceptibility
measurements were made with a Sherwood Scientific balance model MSB-auto. Magnetic moments are
reported at room temperature (298 K) and have been corrected for diamagnetic contributions estimated from
Pascal constants.
[19]
Complex 1 was prepared as described in Ref. 6. Pentafluorophenol, hydrogen chloride
(1.1 M solution in ether), trifluoroacetic acid and triflic acid were purchased from Aldrich and used as
received.
4.2. General procedure for the reaction of 1 with protic acids in stoichiometric ratio 1:1. A solution containing
approx. 440 mg of compound 1 (0.7 mmol) in 40 ml of THF was stirred at -80 ºC, and a second solution
containing exactly the equimolar amount of the corresponding acid (HY = perfluorophenol, 0.46 M in hexane;
trifluoroacetic acid, 0.88 M in toluene; hydrogen chloride, 1.15 M in diethyl ether; triflic acid, 0.47 M in diethyl
ether) was added drop-wise. No colour change was immediately observed, except in the case of triflic acid,
which causes the mixture to turn to a reddish hue. The cooling bath was removed and the mixture was
allowed to warm slowly. With the rest of the acid reagents (perfluoprophenol, hydrogen chloride and
trifluoroacetic acid), colour changes were observed only when the temperature rises to ca. 0 ºC. After
stirring the mixture for 1 h at room temperature, 40 ml of hexane were added. This caused the precipitation
of the product, which was isolated by filtration and dried in vacuum. The remaining solution was taken to
dryness, extracted with hexane (ca. 40 ml) and filtered. The resulting extract has the characteristic purple
colour of 1, and after concentration and cooling to -20 ºC, a small amount of crystals of this compound can
be recovered. The solids were dried under vacuum, affording crude yields of products 2 – 5 that were below
50 % of the initial amount of 1.
8
4.3. General procedure for the reaction of 1 with protic acids in stoichiometric ratio 1:2. This is the same
procedure described above, but double amounts of the HY reagents were used. After precipitation with
hexane the products were filtrated leaving a nearly colourless solution, which evinces total consumption of 1.
The solids were dried under vacuum and recrystallized as indicated below.
4.4. Complex 2: Green crystals. Recrystallized from cold THF (-20 ºC), Isolated yield, 61 % (0.46 mmol from
1
0.75 mmol of 1). H NMR (C6D6, 298 K, 300 MHz), δ (ppm): -22.6 (Δν1/2 = 53, 6H, CH3-C=NAr); 9.8 (Δν1/2 =
107 Hz, 12 H, o-CH3 Aryl); 12.7 (Δν1/2 = 15, 4H, m-CH Aryl); 17.7 (Δν1/2 = 8, 6H, p-CH3 Aryl); 34.8 (Δν1/2 = 27,
-4
1H, H4py); 72.2 (Δν1/2 = 40, 2H, H3py). UV-Vis (CH2Cl2 , C = 10 M): λ(max) = 295 nm (ε = 8200); 350 nm( ε =
2600), 476 nm (ε = 950), 515 nm (ε = 1000), 580 nm (shoulder, ε = 830). IR (Nujol mull): 1641, 1569 ν(C=N,
C=C py) 1259,1214 ν(Car-O); 1014, 995 ν(C-F). µeff (magnetic balance): 5.5 µB. Anal. Calc. for
C39H31F10FeN3O2: C 57.16, H 3.81, N 5.13. Found: C 56.70, H 3.95, N 5.03.
4.5. Complex 3: Burgundy-red solid. Recrystallized from CH2Cl2. Isolated yield, 40 % (0.26 mmol from 0.65
1
mmol of 1). H NMR (CD2Cl2, 298 K, 300 MHz), δ (ppm): –11.2 (Δν1/2 = 49 Hz, 1 H, H4py); 4.1 (Δν1/2 = 62 Hz,
6 H, CH3-C=NAr); 13.8 (Δν1/2 = 107 Hz, 12 H, o-CH3 Aryl); 15.1 (Δν1/2 = 23 Hz, 4 H, m-CH Aryl); 19.2 (Δν1/2 =
19 Hz, 6 H, p-CH3 Aryl); 84.6 (Δν1/2 = 49 Hz, 2 H, H3py). µeff (magnetic balance): 5.13 µB.. UV-Vis (CH2Cl2, C =
-4
10 M): 295 nm (ε = 5500); 350 nm (shoulder, ε = 2500), 625 nm (ε = 540). IR (Nujol mull): 1697 νs(CO2),
1635 ν(C=N), 1588 ν(C=C py), 1260 νas(CO2), 1202 ν(C-F). Anal. Calcd. for C32H33Cl2F6FeN3O4: C 54.80, H
4.60, N 6.18. Found: C 54.56, H 4.67, N 5.54.
4.6. Complex 4.
[3]
Dark blue solid. Washed with 40 ml of hexane and dried under vacuum. Isolated yield, 53
1
% (0.34 mmol from 0.65 mmol of 1). H NMR (CD2Cl2, 298 K, 300 MHz), δ (ppm): -22.2 (Δν1/2 = 86 Hz, 6 H,
CH3-C=NAr); 12.5 (Δν1/2 = 179 Hz, 12 H, o-CH3 Aryl); 15.5 (Δν1/2 = 73 Hz, 4 H, m-CH Aryl); 21.7 (Δν1/2 = 66
Hz, 6 H, p-CH3 Aryl); 35.5 (Δν1/2 = 84 Hz, 1 H, H4py); 82.2 (Δν1/2 = 94 Hz, 2 H, H3py). UV-Vis. (CH2Cl2 , C =
-4
-4
10 M): UV-VIS (CH2Cl2 C = 10 M): 285 nm (ε = 7700), 325 nm (shoulder, ε = 4700), 675 nm (ε = 331). IR
(Nujol mull): 1620 ν (C=N), 1587 ν(C=C py).
4.7. Complex 5. Magenta crystals. Recrystallized from cold CH2Cl2. Isolated yield, 57 % (0.30 mmol from
1
0.52 mmol of 1). H NMR (CD2Cl2 , 298 K, 300 MHz), δ (ppm): -3.6 (Δν1/2 = 52 Hz, 6 H, CH3-C=NAr); 0.8
(Δν1/2 = 133 Hz, 12 H, o-CH3 Aryl); 11.6 (Δν1/2 = 24 Hz, 4 H, m-CH Aryl); 20.3 (Δν1/2 = 11 Hz, 6 H, p-CH3
Aryl); 22.3 (Δν1/2 = 63 Hz, 1 H, H4py); 78.7 (Δν1/2 = 69 Hz, H3py).
(Δν1/2 = 965 Hz).
19
19
F NMR (CD2Cl2, 298 K, 376 MHz): -14.6
-4
F NMR (MeCN, 298 K, 376 MHz): -71.0 (Δν1/2 = 877 Hz). UV-Vis(CH2Cl2 , C = 10 M):
9
λ(max) = 299 nm (ε = 5500); 355 nm (shoulder, ε = 1800); 550 nm (ε = 820). µeff: 4.9 µB (magnetic balance,
25°C). IR (Nujol mull): 3345 ν(OH), 1638, 1619 ν(C=N),1588 ν(C=C py), 1304, 1214, 1165 ν(C-F). Anal.
Calcd. for C29H33F6FeN3O7S2·0.5CH2Cl2: C 43.64, H 4.22, N 5.17. Found, C 43.55, H 4.084, N, 5.36.
4.8. X-ray crystal structure analyses of 2 and 5. A summary of the crystallographic data and the structure
refinement is reported in Table 2. Crystals coated with dry perfluoropolyether were mounted on a glass fiber
and fixed under a cold nitrogen stream. The Intensity data were collected on a Bruker-Nonius X8ApexII CCD
area detector diffractometer using Mo-Kα radiation source (λ = 0.71073 Å) and graphite monochromator. The
data collection strategy used was φ and ω rotations with narrow frames. Instrument and crystal stability were
evaluated from the measurement of equivalent reflections at different measuring times and no decay was
observed. The data were reduced using SAINT
[20]
and corrected for Lorentz and polarisation effects, and a
semiempirical absorption correction was applied using SADABS.
methods using SIR-2002
SHELXTL-6.12
[23]
[22]
[21]
The structures were solved by direct
2
and refined against all F data by full-matrix least-squares techniques using
2
2 2
minimizing w[Fo -Fc ] . All the non-hydrogen atoms were refined with anisotropic
displacement parameters. The hydrogen atoms of compounds 2-5 were included from calculated positions
and allowed to ride on the attached atoms with isotropic temperature factors (Uiso values) fixed at 1.2 times
(1.5 times for methyl groups) those Ueq values of the corresponding attached atoms. Complete structural
data have been deposited with CCDC Reference Nos. 964359 (2) and 964360 (5).
Table 2. Crystallographic data collection, intensity measurements and structure refinement parameters for 2
and 5.
2
5
fw
C39H31F10FeN3O2
•2(C4H8O)
963.73
crystal system
Orthorhombic
Monoclinic
space group
Pnma
P21/n
a, Å
15.7849(5)
15.0793(6)
b, Å
19.3433(6)
8.8111(4)
c, Å
14.5566(5)
25.3883(11)
α, deg.
90.00
90.00
β, deg.
90.00
102.926(2)
γ, deg.
90.00
90.00
4444.6(2)
3287.7(2)
100(2)
4, 1992
100(2)
4, 1584
1.440
1.555
0.429
0.670
θmax, deg
30.5
30.6
no. reflns collected
106012
86096
no.reflns used
6976
10003
formula
V, Å
3
Temperature, K
Z, F(000)
Dcalc, Mgm
µ, mm
-3
-1
10
C29H33F6FeN3O7S2
769.55
no. of param.
2
2 [a]
R1(F) [F >2σ(F )]
2 [b]
wR2(F ) (all data)
[c]
S (all data)
[a]
R1(F) = ∑(
[b]
2
343
441
0.0394
0.0540
0.1151
0.1592
1.071
1.062
Fo - Fc )/∑ Fo
2
2
2
for the observed reflections [F >2σ(F )].
2 2
2 2 1/2 [c]
2
2 2
1/2
wR2(F ) = {∑ [w(Fo - Fc ) ]/ ∑ w(Fo ) } . S = {∑ [w(Fo - Fc ) ]/(n-p)} ;
(n = number of reflections, p = number of parameters).
Acknowledgements
This work was supported by the Government of Spain (project number CTQ2012-30962), the Junta de
Andalucía (project number FQM5074) and the European Union (EU) (FEDER funds). MAC thanks CSIC for
a predoctoral research fellowship (I3P program).
References
[1]
a) L. Li, P. T. Gomes, in Olefin Upgrading Catalysis with Nitrogen-based Metal Complexes, State of
the Art and Perspectives., Vol. II (Eds.: G. Giambastiani, J. Cámpora), Springer, Dordrecht, 2011,
pp. 77-198. b) C. Bianchini, G. Giambastiani, I. Guerrero Rios, G. Montovani, A. Meli, A. M. Segarra,
Coord. Chem. Rev. 2007, 250, 1391-1418.
[2]
M. W. Bouwkamp, E. Lobkowsky, P. J. Chirik, J. Am. Chem. Soc. 2005, 127, 9660-9661.
[3]
G. J. P. Britovsek, M. Bruce, V. C. Gibson, B. S. Kimberley, P. J. Maddox, S. Mastroianni, S. J.
McTavish, C. Redshaw, S. G., S. Strömberg, A. J. P. White, D. J. Williams, J. Am. Chem. Soc. 1999,
121, 8728-8740.
[4]
a) K. P. Bryliakov, N. V. Semikolenova, V. A. Zakharov, K. P. Talsi, Organometallics 2004, 23, 53755378. B) K. P. Bryliakov, E. P. Talsi, N. V. Semikolenova, V. A. Zakharov, Organometallics 2009, 28,
3225-3232.
[5]
a) M. Bochmann, J. Organomet. Chem. 2004, 689, 3982-3998. b) M. Bochmann, Organometallics
2010, 29, 4711-4740.
[6]
J. Cámpora, A. M. Naz, P. Palma, E. Álvarez, Organometallics 2006, 24, 4878-4881.
[7]
pKa values in water. Note that the acid strength of HY is inversely related to the basicity of Y
-
ligands, which correlates with their ability to donate electron density to the metal centre. Aqueous
pKa were selected because they are available for the four acids used in this work, but note that the
acid strengths of HOTf, HCl, acetic acid and phenol holds in water, acetonitrile and dichloroethane,
see ref. [8a].
11
[8]
a) E. Raamat, K. Kaupmees, G. Ovsjannikov, A. Tummal, A. Kütt, J. Saame, I. Koppel, I. Kaljurand,
L. Lipping, T. Rodima, V. Pihl, I. A. Koppel, I. Leito, J. Phys. Org. Chem. 2013, 26, 162-170. b) J.F.J.
Dippy, S.R.C. Hughes, A. Rozanski, J. Chem. Soc. 1959, 2492 - 2498. c) D. Stefanidis, S. Cho, S.
Dhe - Paganon, W. P. Jenks, J. Am. Chem. Soc. 1993, 115, 1650 – 1656.
[9]
I. Fernández, R. J. Trovitch, E. Lobkovsky, P. J. Chirik, Organometallics 2008, 27, 109-118.
[10]
R. J. Trovitch, E. Lobkovsky, M. W. Bouwkamp, P. J. Chirik, J. Am. Chem. Soc. 2009, 27, 62646278.
[11]
W. A. Addison, T. N. Rao, J. Reedijk, J. van Rijn, G. C. Verschoor, J. Chem. Soc., Dalton Trans,
1984, 1349 – 1356.
[12]
G. J. P. Britovsek, J. England, S. K. Spitzmesser, A. J. P. White, D. J. Williams, Dalton Trans. 2005,
945-955.
[13]
Y. M. Baidei, M. A. Siegler, D. P. Goldberg, J. Am. Chem. Soc. 2011, 133, 1274-1277.
[14]
F. A. Cotton, G. Wilkinson, C. A. Murillo, M. Bochmann, Advanced Inorganic Chemistry, 6th ed.,
John Wiley & Sons, New York, 1999.
[15]
K. L. Tokarev, M. A. Kiskin, A. A. Sidorov, G. G. Aleksandrov, V. N. Ikorskii, I. P. Suzdalev, V. N.
Novotortsev, I. L. Eremenko, Russ. Chem. Bull. 2008, 57, 1209-1214.
[16]
D. Lee, S. J. Lippard, Inorg. Chim. Acta 2002, 341, 1-11.
[17]
a) R. C. Mehrotra, R. Bohra, Metal Carboxylates, Academic Press, London, 1983. B) A. Behr, R. He,
K. D. Juszak, C. Krüger, Y. H. Tsay, Chem. Ber. 1986, 119, 991-1015. c) E. Nakamoto, Infrared
Spectra of Inorganic and Coordination Compounds, 4th ed., John Wiley & Sons, New York, 1986.
[18]
G. J. P. Britovsek, V. C. Gibson, S. K. Spitzmesser, K. P. Tellmann, A. J. P. White, D. J. Williams, J.
Chem. Soc. Dalton Trans. 2002, 1159-1171.
[19]
G. A. Bain, J. F. Berry, J. Chem. Ed. 2008, 85, 532-536.
[20]
SAINT, APEX2 Suite, v. 2.1. Bruker AXS Inc., Madison, Wisconsin, USA, 2006.
[21]
G. Sheldrick, SADABS, APEX2 Suite, v. 2.1. Bruker AXS, Inc., Madison, Wisconsin, USA, 2006.
[22]
SIR-2002, APEX2 Suite, v., 2.1, Bruker AXS Inc., Madison, Wisconsin, USA, 2006.
[23]
SHELXTL,v. 6.14. Bruker AXS Inc., Madison, Wisconsin, USA, 2000-2003.
12